Download High-speed and drop-on-demand printing with a pulsed

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Power inverter wikipedia , lookup

Electrical substation wikipedia , lookup

Three-phase electric power wikipedia , lookup

Electrical ballast wikipedia , lookup

History of electric power transmission wikipedia , lookup

Variable-frequency drive wikipedia , lookup

Chirp compression wikipedia , lookup

Islanding wikipedia , lookup

Rectifier wikipedia , lookup

Ohm's law wikipedia , lookup

Current source wikipedia , lookup

Pulse-width modulation wikipedia , lookup

Resistive opto-isolator wikipedia , lookup

Switched-mode power supply wikipedia , lookup

Metadyne wikipedia , lookup

Power electronics wikipedia , lookup

Surge protector wikipedia , lookup

Voltage regulator wikipedia , lookup

Power MOSFET wikipedia , lookup

Buck converter wikipedia , lookup

Stray voltage wikipedia , lookup

Opto-isolator wikipedia , lookup

Voltage optimisation wikipedia , lookup

Alternating current wikipedia , lookup

Rectiverter wikipedia , lookup

Mains electricity wikipedia , lookup

Transcript
Home
Search
Collections
Journals
About
Contact us
My IOPscience
High-speed and drop-on-demand printing with a pulsed electrohydrodynamic jet
This article has been downloaded from IOPscience. Please scroll down to see the full text article.
2010 J. Micromech. Microeng. 20 095026
(http://iopscience.iop.org/0960-1317/20/9/095026)
View the table of contents for this issue, or go to the journal homepage for more
Download details:
IP Address: 75.237.155.60
The article was downloaded on 22/08/2010 at 00:18
Please note that terms and conditions apply.
IOP PUBLISHING
JOURNAL OF MICROMECHANICS AND MICROENGINEERING
doi:10.1088/0960-1317/20/9/095026
J. Micromech. Microeng. 20 (2010) 095026 (8pp)
High-speed and drop-on-demand printing
with a pulsed electrohydrodynamic jet
S Mishra, K L Barton, A G Alleyne, P M Ferreira and J A Rogers
University of Illinois, Urbana, IL 61801, USA
E-mail: [email protected]
Received 5 April 2010, in final form 4 June 2010
Published 20 August 2010
Online at stacks.iop.org/JMM/20/095026
Abstract
We present a pulsed dc voltage printing regime for high-speed, high-resolution and
high-precision electrohydrodynamic jet (E-jet) printing. The voltage pulse peak induces a very
fast E-jetting mode from the nozzle for a short duration, while a baseline dc voltage is picked
to ensure that the meniscus is always deformed to nearly a conical shape but not in a jetting
mode. The duration of the pulse determines the volume of the droplet and therefore the feature
size on the substrate. The droplet deposition rate is controlled by the time interval between
two successive pulses. Through a suitable choice of the pulse width and frequency, a
jet-printing regime with a specified droplet size and droplet spacing can be created. Further, by
properly coordinating the pulsing with positioning commands, high spatial resolution can also
be achieved. We demonstrate high-speed printing capabilities at 1 kHz with drop-on-demand
and registration capabilities with 3–5 μm droplet size for an aqueous ink and 1–2 μm for a
photocurable polymer ink.
(Some figures in this article are in colour only in the electronic version)
become critical. Finally, as with any manufacturing process,
throughput rates (in this case, printing speeds) and process
robustness are key decision parameters in the adoption of the
process. Therefore, to fully realize the capability of the E-jet
printing process, this research reports on how to exploit input
voltage modulation to enhance droplet deposition rates, obtain
consistent droplet volume and accurate spatial placement of
droplets.
E-jet printing uses electric-field induced fluid flows
through fine micro capillary nozzles to create devices in
the micro- and nano-scale range [1]. Typically, these
electric fields are created by establishing a constant voltage
difference between the nozzle carrying the ink (the print
head) and the print substrate. The electric field attracts ions
in the fluid toward the substrate, deforming the meniscus
to a conical shape and eventually leading to instability that
results in droplet release from the apex of the cone [1, 6].
Electrohydrodynamic discharge from a nozzle results naturally
in a pulsed flow. This was exploited by Chen [9] to accurately
place drops. Juraschek and Rollgen [7] reported that this
pulsing persists in the spray regime reporting both lowfrequency 10 Hz and high-frequency 1 kHz pulsations in a
electrohydrodynamic spray. To exploit this natural pulsation,
1. Introduction
Jet printing-based manufacturing processes at the nano- and
micro-scales have been the target of much research because of
the ability to generate very small-scale droplets. Examples of
jet printing include the now ubiquitous ink-jet printing using
thermal and piezo-excitation, and E-jet printing. Among these,
E-jet printing has demonstrated superior resolution, printing
of micron and sub-micron scale droplets using a wide variety
of inks [1–4]. However, the speed of the process and its
ability to produce uniform printing quality have been cited as
impediments, as pointed out in a review on E-jet [5].
Because of the ability to print high-resolution droplets and
lines with a range of inks, E-jet printing has shown tremendous
promise for applications such as printing metallic (Ag)
interconnects for printed electronics [2], bio-sensors [1, 4],
etc. As the advantages of E-jet printing become more apparent
(e.g. the potential for purely additive operations, the ability
to directly print biological materials, maskless lithography),
additional features like drop-on-demand functionality and the
ability to precisely control droplet sizes become necessary.
Further, enhanced process controls to independently regulate
process outputs such as droplet size and delivery frequency
0960-1317/10/095026+08$30.00
1
© 2010 IOP Publishing Ltd
Printed in the UK & the USA
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
Chen et al [9] and Choi et al [8] have developed scaling laws for
characterizing E-jet. Until now, high-resolution E-jet printing
has used this natural pulsation and is therefore limited by the
natural pulsating frequency of the aforementioned discharge,
which has substantial variability. To overcome this limitation,
Kim et al [10] suggested the use of a piezoelectric excitation
of the nozzle tip (hybrid jet printing) along with electric
field induced jetting. Pulsing of ac has been demonstrated
for E-jet by Nyugen et al [11]. Modulation of ac showed
advantages over dc voltage in terms of fabrication of nozzles,
droplet repulsion and drop-on-demand capabilities based on
the frequency of sinusoidal voltage applied. Kim et al [12]
used a square wave dc for E-jet printing and used the amplitude
of the voltage to control the droplet size. Stachewicz et al [13]
demonstrated a single-event pulsed droplet generation for Ejet, as well as a study of relaxation times for drop-on-demand
electrospraying [14].
In the literature cited above, the droplet diameters and
pulse frequencies have been limited to size 50 μm and printing
frequencies of 25 Hz. Further, to the best of our knowledge,
no systematically controlled high-speed printing regimes
have been developed for delivering precise droplet volumes
with high fidelity spatial and temporal resolution. This
paper presents a manufacturing oriented approach to pulsed
input voltage E-jet printing including (1) high-speed printing,
(2) high-resolution printing and (3) a well-documented recipe
for shaping the pulse signal.
In this paper, we present an E-jet printing mode capable
of high speeds and independent control of the droplet size
and printing frequency. Specifically, this mode demonstrates
capability for printing speeds of 1000 droplets per second,
while producing consistent and controllable droplet sizes of
3–6 μm. This mode uses a pulsed voltage signal to generate
electrohydrodynamic flow from the nozzle. The pulse peak
is chosen so as to induce a very fast E-jetting mode from the
nozzle, while the baseline voltage is picked to ensure that a near
conical shaped meniscus is always present, but not discharging
any fluid. The duration of the pulse determines the volume
of the droplet and therefore the feature size on the substrate.
On the other hand, the droplet deposition rate is controlled
by varying the time interval between two successive pulses.
Through suitable choice of the pulse width and frequency, a
jet-printing regime with specified feature size and deposition
rate can be created.
The rest of the paper is organized as follows. Section 2
provides an introduction to the E-jet printing process.
Section 3 then discusses a novel voltage modulation scheme
for delivering high-speed high-resolution E-jet printing
capabilities. A design recipe for determining the parameters
for this scheme is described in section 4. Sections 5 describes
the experimental E-jet printing testbed. Sections 6 and 7
demonstrate high-speed printing and drop-on-demand printing
capabilities of the voltage modulated printing regime. Finally,
in section 8 the contribution of this paper is summarized.
Figure 1. Schematic of the E-jet printing set-up including nozzle
and ink chamber, air supply for back pressure, conducting substrate,
and translation and tilting stages.
Figure 2. Illustration of the change in the meniscus of the fluid due
to an increase in voltage potential between the nozzle tip and the
substrate.
printing include an ink chamber, metal-coated glass nozzle tip,
substrate and positioning system. The controllable printing
process parameters are the back pressure (pneumatic) applied
to the ink chamber, the offset height between the nozzle
and substrate and the applied voltage potential between a
conducting nozzle tip and substrate. Note that the nozzle
tip and substrate are generally coated with metal to ensure
conductivity.
A voltage applied to the nozzle tip causes mobile ions
in the ink to accumulate near the surface at the tip of the
nozzle. The mutual Coulomb repulsion between the ions
introduces a tangential stress on the liquid surface, thereby
deforming the meniscus into a conical shape [1]. At some
point, the electrostatic stress overcomes the surface tension
of the meniscus and droplets eject from the cone. Figure 2
illustrates the change in the ink meniscus due to an increase in
voltage. Depending on the fluid properties, as the applied field
is increased this discharge begins as a pulsed or intermittent jet
(pre-jet modes) transitioning into a stable single jet, multiple
unstable jets and finally becoming a spray for very large
electric field strengths. Each of the different jetting modes
(e.g. pulsating, stable jet, E-spray [15]) can be used to achieve
various printing/spraying applications. Pre-jet modes are
typically used for printing because of better controllability
at high speeds.
Changes in back pressure, stand-off height and applied
voltage affect the size and frequency of the droplets. This
sensitivity of the process output to variations under the printing
conditions requires high-resolution sensing and control in
order to achieve stable and predictable printing results.
2. Electrohydrodynamic jet printing
Figure 1 presents a schematic of the E-jet printing process.
As can be seen from figure 1, the main elements for E-jet
2
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
3. Voltage modulation in E-jet printing
Typically, the jet frequency and droplet diameter have been
controlled by changing the applied voltage difference across
the tip and the substrate. From a process development point
of view, this has significant disadvantages. First, for a given
nozzle diameter, printing ink and stand-off height (distance of
the nozzle tip from the substrate), the droplet diameter on the
surface (D) and jetting frequency (f ) are coupled. Scaling
laws from Choi et al [8] capture this dependence with the
following equations:
√
3
E2
E
f =
and
d=
(1)
dN
dN
D = dF (θ ),
(2)
Figure 3. Schematic time plot of the voltage profile for the pulsed
E-jet. Td denotes the time between successive pulses while Tp
denotes the pulse width. Vh and Vl are the high and low voltages
respectively.
where dN is the anchoring radius of the meniscus, d is the
droplet diameter of the ejected droplet, E is the electric field
because of the applied potential and θ is the contact angle at
the surface; F (θ ) is a function of the contact angle θ . As
can be seen from the above equations, one can set a voltage
level to either obtain a desired droplet diameter or a printing
speed (droplets/sec), but not both. The second disadvantage
associated with printing by setting a constant voltage different
between the tip and substrate accrues from the fact that minute
changes in the stand-off height (for example, because of small
misalignments or errors associated with the motion stage) can
cause significant changes in the jetting frequency and droplet
diameters.
With a sufficiently high potential difference, very fast
jetting frequencies of several kHz can be achieved. However,
at the resulting strong electric field, the system becomes
more sensitive to variations under operating conditions such
as stand-off height, meniscus wetting properties, etc and
the jetting frequency may vary substantially during printing,
leading to inconsistent droplet spacing. Therefore, constant
high-voltage E-jetting is unsuitable for printing large droplet
arrays with regular droplet diameters and consistent droplet
spacing (as might be required in a DNA microarray, for
example). At the same time, low-voltage E-jetting results
in slow printing speeds (with droplet deposition rates of 1–
5 drops per second).
In order to be able to overcome these limitations, we
propose a short-time high-voltage pulse superimposed over a
lower baseline constant voltage. The short high-voltage pulse
releases a droplet (or a finite number of droplets) from the
nozzle, while the lower constant voltage holds the charge in
the meniscus. Figure 3 shows the time plot of a typical
pulse. The duration of the pulse controls the number
of droplets released. These droplets coalesce and form a
larger droplet on the substrate surface. Hence the volume
of fluid deposited on the substrate is controlled by the number
of droplets released per high-voltage pulse and consequently,
the duration of this pulse. On the other hand, the time between
two pulses determines the time or (for constant velocity motion
of the stage) distance (s) between successive droplets on the
substrate.
The baseline voltage must be chosen such that there is
no jetting at that voltage; however, it must be large enough
to ensure that the Taylor cone [16] is formed and maintained
at the tip of the micro capillary nozzle. On the other hand,
the pulse peak voltage Vh is chosen such that it results in a
very fast natural jetting mode with a frequency of jetting given
by fh . By adjusting the pulse peak voltage to a large enough
value, it is possible to get fh of the order of 10–50 kHz for
most inks [1, 9].
3.1. Pulse spacing Td
The pulse spacing Td directly controls the droplet spacing on
the substrate. This is because the distance between droplets
can be changed by adjusting the time between successive
pulses and the speed of movement (wst ) of the substrate with
respect to the nozzle tip. The droplet spacing is given by
sd = wst Td .
3.2. Pulse width Tp
Assuming a hemispherical droplet of D on the surface of the
substrate, we have (for fh Tp > 2)
fh vh Tp =
π 3
D ,
12
(3)
where vh is the volume of a single droplet released from the
nozzle and Tp is the pulse width. Given a fixed Vh (pulse peak),
fh and vh are fixed. Therefore we can control the diameter of
the deposited droplets by changing the pulse width Tp . Further,
the size of these ‘aggregated’ droplets is more uniform than
each individual discharged droplet because of the averaging
effect. For a small enough pulse width, there may be no droplet
released from the tip because of the time delay in formation
of the meniscus and ejection of the droplet. This minimum
possible Tp is dependent on the choice of Vh (see figure 13 for
an example of a recorded input voltage pulse and the resulting
current signal). Figure 4 shows a plot of this relationship
for a photocurable polyurethane polymer (Norland Optical
Adhesive NOA 73).
Therefore, by adjusting Tp and Td we can fix the desired
droplet diameter and spacing independently.
3
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
Figure 4. Plot of the minimum pulse width Tp against the input
voltage Vh for a polyurethane polymer ink. For larger voltages, we
can obtain droplet ejection for smaller pulse widths. For Vh =
425 V, we obtain fh > 18 kHz.
Figure 5. 5-axis experimental testbed for E-jet printing.
4. Design recipe
Table 1. System components.
In this section, we algorithmically describe how the input
parameters, specifically the pulse modulation parameters
Vh , Vl , Tp and Td are determined, based on output
requirements of the printing process, such as droplet spacing
and droplet (feature) size.
(i) Set process parameters: ink type, substrate type, back
pressure (psi), and nozzle diameter. Typically, the nozzle
diameter is chosen to be between 2 and 5 times the desired
droplet diameter, while the back pressure is chosen so that
it holds a spherical meniscus at the nozzle tip.
(ii) Set Vl to be 5–10 V less than the initial voltage required to
release a single droplet. (This voltage to release a droplet
can be arrived at by gradually raising the voltage until the
first drop is released from the nozzle.)
(iii) Determine the substrate velocity (fastest possible, subject
to motion stage hardware constraints) wst .
(iv) Determine the time between pulses as Td = wsdst , where sd
is the desired spacing between droplets.
(v) Evaluate the potential Vh range so that
(a) Vh is significantly less than spraying, unstable jetting
or arcing voltage.
(b) Vh is greater than the voltage that results in a jetting
frequency of fh,min which satisfies fh,min Td > 2.
(vi) Choose Vh to be as large as possible without violating (a)
above. Determine the pulse width Tp to ensure the desired
droplet diameter D, based on equation (2).
Part
Manufacturer
Resolution
X, Y, Z stages
Infinity 3 camera
Zoom lens
Illuminator
Voltage amplifier
Current detector
Aerotech
Lumenera
EdmundOptics NT55-834
EdmundOptics NT55-718
Trek 677B
Femto NT59-178
0.01 μm
2 Mpixel
2.5×–10×
N/A
1V
1 nA
alignment and jetting visualization. The translational motion
of the substrate is controlled through the X, Y axes, while
the pitch and yaw are fixed through the U, A axes. The
electrical connection to the nozzle and substrate, along with
the substrate-side measurement scheme, followed the set-up
illustrated in figure 10. The measured current signal was
detected online and processed off-line to determine jetting
information. The printing was done on a glass slide substrate
coated with Au for conductivity. No other post-processing of
the substrate was done. The standoff height for printing was
set at 30 μm along the Z axis. For the interested reader, the
effect of the stand-off height has been investigated in [8]. The
power supply used is bipolar; however, for this paper, we will
use positive polarity of the nozzle for demonstrating printing.
We now present printing results for (a) high-speed printing, and (b) drop-on-demand printing in the following sections.
6. High-speed printing results
The pulsed E-jet printing can significantly enhance printing
speeds (droplet deposition rate). Typically in constant jet
mode printing applications, jetting frequencies are around 1–
5 droplets per second [1]. A graphics art rendered pattern
1.5 mm by 0.3 mm was used as a basis for comparison of
printing speeds for constant voltage and pulsed voltage E-jet
printing. The constant voltage printing was executed at 1
droplet per second jetting frequency and required ≈ 2200 s
for printing. On the other hand, using pulsed jetting at
60 droplets per second the pattern was printed in 70 s.
5. System description
To validate the feasibility of both the high-speed and drop-ondemand E-jet printing process, the design scheme described
in the previous section was implemented on an experimental
E-jet printing testbed shown in figure 5. Table 1 describes the
hardware components of the system.
As can be seen from figure 5, the motion control system
consists of five physically connected axes (X, Y, Z, U, A), a
substrate mount, a nozzle mount and a camera for nozzle
4
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
Figure 6. Chart showing printing times for 1.5 mm by 0.3 mm
pattern using a constant voltage jet-printing mode and pulsed
voltage printing jet mode. Pulsed voltage printing required 70 s,
while constant voltage jet printing required 2200 s.
Figure 8. Printed pattern using NOA 73 (photocurable polyurethane
polymer) at 1 kHz printing frequency using a 2 μm ID capillary
nozzle. The droplet diameter varied from 1 to 2 μm.
(a)
Further, pulsed E-jetting shows tremendous potential for
establishing printing speeds well into several kHz that will
transform this technology into a viable nano-manufacturing
process. Figure 8 illustrates an image printed at 1 kHz with
droplet sizes ranging from 1 to 2 μm. Printed lines can be
laid down on a substrate at speeds up to 10 kHz, shown in
figure 9.
(b)
Figure 7. (a) Printed pattern using constant voltage jetting (5 μm
capillary, phosphate buffer solution with 10% glycerol (vol.)). Total
area = 0.3 mm × 1.5 mm. (b) Printed pattern using pulsed voltage
jetting (5 μm capillary, phosphate buffer solution with 10% glycerol
(vol.)) 0.3 mm × 1.5 mm. Printing with constant jetting resulted in
irregular droplet spacing and size and required 2200 s. Printing with
pulsed jetting resulted in regular droplet spacing, consistent droplet
sizes and was completed in 70 s. Typical droplet diameter was 3 μm.
7. Drop-on-demand printing
7.1. Current detection
Figure 7 shows optical micrographs of the printed patterns
obtained from the two printing methods. As shown in
the chart in figure 6, the printing time was cut down by
a factor of 30 using the pulsed mode operation. The
droplet placement accuracy using the pulsed mode operation
is critically dependent on the synchronization of the stage
movement and the voltage pulsing. This is the source of
the irregular droplet alignment from one raster to the next in
figure 7.
Monitoring the E-jet process optically becomes very
challenging especially with printing at single micron
resolutions and speeds approaching 1000 droplets per second.
Therefore, a scheme based on current monitoring is developed
for the process. Essentially, the E-jet process involves
combined mass and charge transfer between the nozzle and
substrate, i.e. each droplet released from the nozzle carries a
net positive or negative charge [3], depending on the direction
5
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
Figure 11. Illustration of the one-to-one correlation between the
printed droplets and the measured current peaks.
the substrate-ground connection. This measurement scheme
is termed substrate-side measurement. Figure 10 shows a
schematic of substrate-side current measurement set-up. The
high-voltage source is connected to the nozzle side, while a
current sensor is connected to the substrate side. The free end
of the current sensor drains to ground. While both schemes
work well for process monitoring, we will use the substrateside configuration for the experimental results presented
below.
The frequency of jetting can be determined by measuring
the time elapsed between two successive jets. Each peak in the
current signal corresponds to a single jet. This is illustrated
in figure 11. For the resolution range (<5 μm) in which
we operate, the typical measured current associated with each
droplet is found to be in the range of 10 to 100 nA. Thus,
the jet current detection capability, while not necessary for
pulsed mode E-jetting, is useful for determining the number
of droplets released per pulse. This information can then
be used for establishing voltage pulse modulation parameters
described in sections 3 and 4.
Figure 9. SEM images of printed lines using NOA 73 (photocurable
polyurethane polymer) at 10 kHz printing frequency using a 2 μm
ID capillary nozzle. The zoomed-in detail on the right shows the
spreading of the droplets after printing.
7.2. Printing results
We demonstrate additional capabilities of the high-speed
pulsed E-jet printing regime through the following. Figure 12
shows a time plot of current measurement on the substrate
side superimposed on a time plot of the voltage pulse. The
ink used was a 10 mM aqueous phosphate buffer solution with
10% (by vol.) glycerol. We observe that a single droplet is
released from the micro capillary within the pulse time. This
capability directly translates into a drop-on-demand regime
for E-jet, which will substantially enhance the applicability of
E-jet for printing bio-sensors, among other applications.
Figure 13 shows the measured current plot for a train of
voltage pulses and the corresponding droplet ejections. We
note that multiple droplets (in this case, four) are ejected
within each pulse period. By changing the pulse time (Tp ),
the number of droplets ejected per pulse can be controlled.
Figure 10. Schematic of the substrate-side current measurement
set-up for the E-jet process.
of the applied field. With the release of each charged droplet
from the nozzle, a small current is drawn to neutralize the
resulting charge imbalance in the fluid at the meniscus. By
detecting this current, the time of droplet release can be
determined. This measurement scheme is termed nozzleside measurement. An alternate scheme measures the current
dissipated through the substrate to ground when a charged
droplet from the nozzle arrives at a conductive substrate. This
current can be measured by introducing a current sensor in
6
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
Figure 12. Plot of current measurement showing a voltage pulse
and the corresponding peak of a single droplet.
(b)
(a)
Figure 15. (a) Printed pattern using NOA 73 from a 5 μm micro
capillary, with on-the-fly droplet diameter control by changing the
pulse width Tp . (b) Detail of pattern showing controlled transition
from 3.9 to 8.1 μm droplet size. The droplet size is controlled
independent of droplet spacing (16 μm).
through using a micro capillary nozzle of inner diameter (ID)
5 μm. The droplet diameter varied from 6 to 22 μm based on
the duration of the pulse width.
The pulsed mode operation of the E-jet process enables
on-the-fly droplet diameter change. This is illustrated in
figure 15. The droplet diameter is varied by changing the
pulse width and generates denser and less dense printed areas
without the need for changing nozzle tips, readjustment of
voltage or change in deposition frequency. We therefore
independently control droplet diameter and droplet spacing,
as mentioned in section 3.
This independent control of droplet spacing and droplet
diameter can be exploited to create patterns with varying
density of droplets or varying droplet size that can be adjusted
on the fly. Figure 15 demonstrates this capability in a printed
pattern using NOA 73 from a 5 μm micro capillary. The
Figure 13. Plot of current measurement showing a voltage pulse
train with multiple droplets released per pulse.
Through varying pulse time, multiple droplets can jet
within the pulse width and coalesce to create different droplet
sizes. Figure 14 shows a plot of varying pulse width (Tp )
against droplet diameter (D) on the substrate. The ink was
a UV curable polyurethane ink; jetting was accomplished
Figure 14. Plot of droplet diameter on the surface (D) against the pulse width Tp . The predicted slope of 0.33 is plotted as a dashed line.
We see good correlation between the prediction and the measurement values.
7
J. Micromech. Microeng. 20 (2010) 095026
S Mishra et al
droplet size is varied by changing the pulse width (Tp ) from
500 to 2500 μs. The resulting average droplet size is found
to be 3.9 μm and 8.1 μm respectively for the two cases, with
standard deviations of 0.4 μm and 0.3 μm (with 16 random
droplet diameter measurements). The droplet spacing (16 μm)
is unaffected by changes in droplet size.
References
[1] Park J-U et al 2007 Nature Mater. 6 782–9
[2] Lee D-Y, Lee J C, Shin Y-S, Park S-E, Yu T-U, Kim Y-J
and Hwang J 2008 J. Phys.: Conf. Ser. 142 012039
[3] Park J-U, Lee S, Unarunotai S, Sun Y, Dunham S, Song T,
Ferreira P M, Alleyne A G, Paik U and Rogers J A 2010
Nano Lett. 2 584–91
[4] Park J-U, Lee J H, Paik U, Lu Y and Rogers J A 2008 Nano
Lett. 8 4210–6
[5] http://technologyreview.com/computing/19373/page1/
[6] Jaworek A and Krupa A 1996 J. Aerosol Sci.
27 979–86
[7] Juraschek R and Rollgen F W 1998 Int. J. Mass Spectrom.
177 1–15
[8] Choi H K, Park J-U, Park O O, Ferreira P M,
Georgiadis J G and Rogers J A 2008 Appl. Phys. Lett.
92 123109
[9] Chen C H, Saville D A and Aksay I A 2006 Appl. Phys. Lett.
89 124103(1)–(3)
[10] Kim Y-J, Kim S-Y, Lee J-S, Hwang J and Kim Y-J 2009
J. Micromech. Microeng. 19 107001
[11] Nguyen V D and Byun D 2009 Appl. Phys. Lett. 94 173509
[12] Kim J, Oh H and Kim S-S 2008 J. Aerosol Sci.
39 819–25
[13] Stachewicz U, Yurteri C U, Marijnissen J C M and
Dijksman J F 2009 Appl. Phys. Lett.
95 224105
[14] Stachewicz U, Dijksman J F, Burdinski D, Yurteri C U
and Marijnissen J C M 2009 Langmuir
25 2540–9
[15] Cloupeau M and Prunet-Foch B 1994 J. Aerosol Sci.
25 1021–36
[16] Taylor G 1969 Proc. R. Soc. Lond. A. 313 453–75
8. Conclusions
E-jet printing technology has shown tremendous potential
for applications in printed electronics, biotechnology and
micro-electromechanical devices. Printing speed and droplet
size control present the biggest challenge for jet-printing
techniques. In order to address these issues simultaneously,
a high-speed high-precision E-jet printing technique was
investigated. By using a pulsed dc voltage signal to produce
E-jetting, precise droplet placement and droplet spacing was
obtained at very fast printing speeds. The printing times
were cut down by three orders of magnitude, while delivering
specified droplet deposition rates and feature sizes. Further,
the proposed method also demonstrated drop-on-demand
capability, as well as on-the-fly droplet feature size and droplet
volume control.
Acknowledgments
The authors gratefully acknowledge the contribution and
support of the NSF Nano-CEMMS Center under award
number DMI-0328162 and CMMI-0749028.
8